N,O-Nucleosides from Ene Reactions of Nitrosocarbonyl Intermediates

Dec 17, 2012 - Flow system does the dirty work. Organic chemists optimizing a reaction, like chefs perfecting a dish, execute a single transformation...
0 downloads 0 Views 1MB Size
Article pubs.acs.org/joc

N,O‑Nucleosides from Ene Reactions of Nitrosocarbonyl Intermediates with the 3‑Methyl-2-buten-1-ol Paolo Quadrelli,* Mariella Mella, Serena Carosso, and Bruna Bovio Dipartimento di Chimica, Università degli Studi di Pavia, Viale Taramelli 12, 27100, Pavia, Italy S Supporting Information *

ABSTRACT: Nitrosocarbonyl intermediates undergo ene reactions with allylic alcohols, affording regioisomeric adducts in fair yields. Nitrosocarbonyl benzene reacts with 3-methyl-2-buten-1-ol and follows a Markovnikov orientation and abstracts preferentially the twix hydrogens over the lone ones. With the more sterically demanding nitrosocarbonyl mesitylene and anthracene, the Markovnikov directing effect is relieved and lone abstraction is observed, affording the 5-hydroxy-isoxazolidines that serve as synthons for the preparation of N,O-nucleoside analogues according to the Vorbrüggen protocol.



INTRODUCTION The discovery of several new series of nucleosides analogues with antiviral activity altered the classical way of thinking about nucleosides analogues as antiviral agents.1 These derivatives play a fundamental role in viral chemotherapy and modification of the sugar fragment and its replacement with a carbocyclic moiety resulted in the syntheses of interesting nucleosides analogues that have shown remarkable activity toward a variety of viruses.2 Methods for the synthesis of sugar- and sugar-modified and carbocyclic-nucleosides were studied extensively.3 However, various synthetic problems were frequently encountered, such as low yields, low stereoselectivity, and toxicity problems of the obtained compounds, which are just some of the problems that characterize these synthetic approaches. Among the various synthetic strategies, the heteroatom substitution produced a series of compounds where the sugar moiety was replaced by alternative heterocyclic rings. Furthermore, the introduction of a side chain on the sugar led to several branched nucleosides that were found to be potential antitumoral or antiviral agents.4 In this field, the Chiacchio and Romeo groups have described the synthesis of 4′-α-C-branched N,O-nucleosides based on the 1,3-dipolar cycloadditions of nitrones with vinyl acetate, followed by coupling with silylated nucleobases. The obtained compounds were evaluated for their activity against various viruses as well as cytoxicity and apoptotic activity.5 © 2012 American Chemical Society

Recently, we have proposed the synthesis of a new class of carbocyclic nucleosides starting from cyclopentadiene using the nitrosocarbonyl intermediates (RCONO, 1) chemistry.6 These intermediates, generated by the mild oxidation of nitrile oxides with tertiary amine N-oxides or by oxidation of hydroxamic acids, are efficiently trapped by cyclopentadiene (Scheme 1) to afford the hetero Diels−Alder (HDA) cycloadducts 2 that are highly reactivity dipolarophiles and were employed to synthesize the conformationally restricted carbocyclic moiety aminols 4 through amide hydrolysis and N−O bond cleavage of the cycloadducts 3.7 Aminols 4 were useful for the linear construction of purine and pyrimidine nucleosides 5. These nucleosides are characterized by the presence of a secondary hydroxy group on the cyclopentane moiety and by an isoxazoline ring fused to the carbocyclic unit and were found moderately active against human Herpes (types 1 and 2) and Varicella viruses.8 Recently, anthryl derivatives were found active with no cellular toxicity at the dose tested against the Human Papilloma virus (HPV).9 Applications of nitrosocarbonyls to the syntheses of biologically active molecules are well-known, essentially due, first, to the variety of generating methods available (oxidation of hydroxamic acids10 and nitrile oxides11 and thermal12 or photochemical13 cycloreversions of 1,2,4-oxadiazole-4-oxides) Received: October 29, 2012 Published: December 17, 2012 516

dx.doi.org/10.1021/jo302346a | J. Org. Chem. 2013, 78, 516−526

The Journal of Organic Chemistry

Article

Scheme 1. Synthetic Pathway of Isoxazoline-Carbocyclic Nucleosides through Nitrosocarbonyl Chemistry and Linear Construction of the Heterobases

Scheme 2. Ene Reactions of Aromatic Nitrosocarbonyl Intermediates with Trisubstituted Olefins

Scheme 3. Ene Reactions of Aromatic Nitrosocarbonyl Intermediates with the 3-Methyl-2-buten-1-ol and Its Methyl Ether and Acetyl Derivatives

and, second, to the exceptional reactivity of these intermediates in HDA reactions. Nevertheless, nitrocarbonyl intermediates are also powerful enophiles,14 but no applications of ene reactions toward the synthesis of nucleosides are actually reported.15 We have detailed that nitrosocarbonyl intermediates, generated at r.t. by the mild oxidation of nitrile oxides, undergo clean ene reactions with trisubstituted olefins 9. Allylic

hydrogens on the more congested side of the alkene are exclusively abstracted (the “cis effect”), thus resembling the singlet oxygen behavior. Nitrosocarbonyl benzene 8P (P = phenyl) follows a Markovnikov (M) orientation and abstracts preferentially the twix hydrogens over the lone ones. With the more sterically demanding nitrosocarbonyl mesitylene 8M (M 517

dx.doi.org/10.1021/jo302346a | J. Org. Chem. 2013, 78, 516−526

The Journal of Organic Chemistry

Article

addition of the nitrosocarbonyl benzene 8P to the lesssubstituted carbon atom of the double bond, in full accordance with a M orientation.14,16 Adduct 13Pd derives from 10Pd simply upon water elimination. When the hydroxy group of the alcohol 9d is protected as an ether in 9e or an ester in 9f, the ene reactions with the nitrosocarbonyl benzene 8P proceed straight to single ene adducts 10Pe and 10Pf, respectively, which are obtained in good yields according to the addition of the nitrosocarbonyl electrophilic nitrogen to the alkene in an M fashion.19 Moving to the nitrosocarbonyl mesitylene 8M case, the experimental conditions were even simpler and cleaner. Mesitonitrile oxide 7 was added to a stirred solution of NMO (1.1 equiv) in DCM in the presence of an excess (5 equiv) of 3-methyl-2-buten-1-ol 9d at r.t. After stirring overnight and column chromatographic separation, the ene adduct 10Md was isolated in 50% yield along with an unexpected second compound, 11Md, in only, but remarkable, 25% yield. Compound 10Md is the M adduct, and its structure is in accordance with the analytical and spectroscopic data. The structure of 11Md was not immediately clear. The 1H NMR spectrum indicates the presence of one deshielded proton at 5.44 δ coupled with two protons at 2.14 and 2.35 δ, while an OH group gives a singlet at 6.98 δ, as confirmed by the corresponding IR spectrum. The 13C NMR spectrum presents a signal at 95.3 δ resembling an acetal-type carbon atom. The structure assignment as shown in Scheme 3 came from the Xray analysis (see the Supporting Information). These types of 5hydroxy-isoxazolidines are usually prepared by reaction of hydroxamic acids and α,β-unsaturated aldehydes.20 The origin of the isoxazolidine 11Md was undoubtedly attributed to a relief in the M addition determined by the more sterically demanding mesityl group that activates the AM path.16 As the addition of the nitrosocarbonyl moiety occurs on the more substituted carbon atom of the allylic alcohol 9d, the primary adduct is the enol 14 that evolves into the nonisolable aldehyde 15, which undergoes cyclization to the hemiacetal 11Md (Scheme 4). In fact, when the protected alcohols 9e,f were used and the reactions were carried out in the same conditions, adducts 10Me and 10Mf were obtained in 63% and 56% yields, respectively, along with the enol ethers 12Me and 12Mf in 35% and 32% yields, respectively (Scheme 3 and Table 1). The structures of the adducts obtained from alkenes 9e,f rely upon their analytical and spectroscopic data. The 1H NMR spectra of adducts 10Me and 10Mf show the olefinic methylenes as singlets at 4.90 and 5.00 δ for 10Me and at 4.96 and 5.04 δ for 10Mf. The N-OH singlets are found at 9.07

= mesityl), the M-directing effect is relieved and comparable twix and lone abstraction are observed (Scheme 2).16 On pursuing our research in nitrosocarbonyl ene reactions, we extended the studies to other trisubstituted olefins and, in particular, to allylic alcohols (R = OH). The 3-methyl-2-buten1-ol was found to be the proper model to investigate the selectivity outcome in the ene reaction with nitrosocarbonyls bearing more sterically demanding substituents. An unexpected isoxazolidine deriving from the anti-Markovnikov (AM) addition to the alcohol is presented as a valuable synthon toward the preparation of new N,O-nucleoside analogues through the Vonbrüggen protocol.



RESULTS Ene Reactions with Nitrosocarbonyl Benzene and Mesitylene. Addition of a dichloromethane (DCM) solution of benzohydroximoyl chloride 6 to a stirred solution of Nmethyl-morpholine N-oxide (NMO, 1.1 equiv) and Et3N (1 equiv or a catalytic amount) in DCM in the presence of an excess (5 equiv) of 3-methyl-2-buten-1-ol 9d afforded, after standing overnight at r.t., the ene adducts 10Pd and 13Pd (Scheme 3) that were isolated upon chromatographic separation of the reaction mixture in the reported yields (Table 1). The same procedure was followed for the allylic ether 9e17 and acetate 9f,18 which were prepared according to the reported syntheses. Table 1. Yields of Adducts in the Reactions of Nitrosocarbonyls 8P,M with Alkenes 9d−9f Ar-CONO

alkene

10

8P

9d 9e 9f 9d 9e 9f

50 70 60 50 63 56

8M

11

12

13 8

25 35 32

The structures of all isolated ene adducts rely upon the corresponding analytical and spectroscopic data. In the 1H NMR spectra, the typical olefinic methylene signals are found in the range between 5.00 and 5.43 δ for adducts 10Pd−10Pf and more shielded between 4.89 and 5.00 δ for the derivative 13Pd. The N-OH group is evident when the spectrum is recorded in DMSO and is found strongly deshielded in the range 8.2−8.4 δ for 10Pd−10Pf and even more (9.3−9.5 δ) for adduct 13Pd. In the reaction with the allylic alcohol 9d, compound 13Pd is related to 10Pd and both derive from the Scheme 4. AM Pathway: The Formation of Isoxazolidine Ring

518

dx.doi.org/10.1021/jo302346a | J. Org. Chem. 2013, 78, 516−526

The Journal of Organic Chemistry

Article

Scheme 5. Ene Reaction of Anthryl Nitrosocarbonyl Intermediate with the 3-Methyl-2-buten-1-ol

Scheme 6. Synthesis of Isoxazolidine-Nucleoside Analogues through the Vorbrüggen Protocol

and 9.27 δ for the two products, respectively. The enol-ether 12Me showed the double bond signals at 5.23 and 6.52 δ and the N-OH singlet at 9.10 δ while the enol-acetate 12Mf showed the double bond signals at 5.88 and 7.21 δ and the N-OH singlet at 9.38 δ. The results of the reactions of nitrosocarbonyls 8P,M and olefins 9d−9f seem to show that hydrogen bonding is ineffective with respect to the selectivity. To verify experimentally the influence of the substitution pattern on the reactivity of the double bond, we performed control experiments by means of intermolecular competition reactions of the alkenes 9d versus 9e and 9f, respectively, in the presence of the nitrosocarbonyl benzene 8P and mesitylene 8M. The results showed that the expected compounds (as shown in Table 1) from the competitive reactions were isolated in all the

cases with the same ratios (see the Experimental Section). The allylic alcohol 9d was not found somehow more reactive than the others, and the selectiviy outcomes were maintained. Enforcing the AM Path. We explored the ene reaction of another typical stable nitrile oxide,21 the anthracenenitrile oxide 17 obtained from the 9-anthraldeyde oxime 16 according to the preparation reported in the literature22 by treating a chloroform solution of 16 with NCS and catalytic pyridine at 0 °C for 2 h. The solid nitrile oxide 17 was added portionwise to a stirred DCM solution of 5 equiv of the allylic alcohol 9d in the presence of 2 equiv of NMO at r.t. for 48 h. After the usual workup, the residue was submitted to column chromatography, and two products were isolated and identified as compounds 19 and 20 in 27% and 52% yields, respectively (Scheme 5). 519

dx.doi.org/10.1021/jo302346a | J. Org. Chem. 2013, 78, 516−526

The Journal of Organic Chemistry

Article

easily detectable. In the 1H NMR spectrum of the uracil derivative 22Aa, the doublets (J = 8 Hz) at 5.49 and 7.86 δ are clearly related to the uracil double bond while the imide NH is found highly deshielded at 11.22 δ as a singlet. In the 1H NMR spectrum of the cytosine derivative 22Mg, the double bond doublets (J = 8 Hz) are found at 5.53 and 7.39 δ. Again, when, in the uracil ring, the CH−C proton is replaced by a halogen (b, c, d, e) or by a methyl, as in the case of tymine (f), a single proton corresponding to the N−CH is observed in the range of 7.29−7.86 δ, while the imide NH singlets remain strongly deshielded in the range of 11.23−11.75 δ. Nucleoside analogues 22M,A can be submitted for biological evaluation without any further structural modification, either for their antiviral behavior or, in other cases, in order to compare their activities with those of reported carbocyclic and heterocyclic structures.5,8,24,28 In particular, in nucleoside analogues 22A, the presence of an anthracene ring adds fluorescent properties to these compounds, a further tool for our future studies. The use of fluorescent moieties in the nucleoside synthesis has attracted the interest of various research groups as marker molecules or in the field of imaging to follow their path inside the cells in order to achieve a better understanding of the “in vivo” mechanisms.29 Fluorescent structures can be considered “molecular labels”30 to be assembled in specific oligofluorosides, as sensors31 or indicators of the behavior of the DNA.32 New fluorophores are constantly developed as well as their applications.33 Fluorescent polyaromatic groups may be active through their ability to establish π-stacking interactions with themselves34 as well as DNA intercalators;35 these mechanisms are in action in different cases depending upon the biological targets.

Their structures rely upon the relative spectroscopic data. In particular, the 1H NMR spectrum of the major compound 20 showed a deshielded proton as a triplet at 5.22 δ, coupled with two protons at 2.20 and 2.44 δ. Just before the aromatic signals (7.5−8.6 δ), a doublet was found at 6.84 δ corresponding to the OH group (exchange with D2O) and whose presence was confirmed by the IR spectrum (νOH = 3200 cm−1). The 13C NMR spectrum showed significantly the acetal-type carbon at 95.8 δ. The experiment performed with anthracene nitrosocarbonyl 18 and alcohol 9d confirms the relief of the M control observed in the addition of 8M to enes 9d−9f (Table 1), and the results show the remarkable influence of the nitrosocarbonyl substituent on the selectivities. The approach of the bulkier anthracene group determines a strong steric hindrance that disfavors the M path enforcing the AM selectivity. Synthesis of N,O-Nucleosides. The 5-hydroxy-isoxazolidines 11Bd and 20 obtained from the ene reaction of nitrosocarbonyl mesitylene 8M and nitrosocarbonyl anthracene 18 with the allylic alcohol 9d were used as synthons to prepare N,O-nucleoside analogues by adapting the Vorbrü ggen protocol23 as well as the Chiacchio’s procedures for similar compounds.5,24 The acetyl derivatives of 21M,A were prepared according to standard procedures (Scheme 6).25 The acetylated compound was obtained in nearly quantitative yield and fully characterized. In the 1H NMR spectrum of 21M, the presence of the acetyl group is shown by the presence of a singlet at 2.09 δ while the OH signal, previously found at 6.98 δ, disappeared. Similarly, in the 1H NMR spectrum of 21A, the acetyl group is found as a singlet at 2.15 δ while the OH signal at 6.84 δ is now absent. The Vorbrüggen protocol can be applied on both previously silylated heterobases or commercial compounds in the presence of silylating agents.23,26 In the cases at hand, we report the syntheses conducted on commercial heterobases of the uracil family used in the presence of the in situ silylating agent, the bis(trimethylsilyl)acetamide (BSA), and trimethylsilyl trifluoromethanesulfonate (TMSO-Tf) as reaction promoter.27 The acetylated isoxazolidine 21M,A were added under a nitrogen atmosphere at r.t. to a solution of the selected heterobases (2 equiv) and BSA (2 equiv), and the solutions became clear after boiling in DCM for a couple of hours. The mixtures were then ice-cooled at 0 °C, TMSO-Tf was added, and the reactions were refluxed overnight (Scheme 6). The desired compounds 22M,A(a−g) were obtained as white solids by simple crystallization or column chromatographic purification. Nucleoside analogues 22M,A(a−g) were isolated from good to excellent yields (48−99%), and their structures rely upon the analytical and spectroscopic data. All the reactions gave single products belonging to the family of uracil derivatives. The uracil derivative 22Ma is representative of a general trend in the 1H NMR spectra of this type of compound, and the doublets (J = 8 Hz) at 5.67 and 7.64 δ are clearly related to the uracil double bond while the imide NH is found highly deshielded at 10.10 δ as a singlet. The same features were found in cytosine derivative 22Mg. When, in the uracil ring, the  CH−C proton is replaced by a halogen (b, c, d, e) or by a methyl, as in the case of tymine (f), a single proton corresponding to the N−CH is observed in the range of 7.43−7.80 δ, while the imide NH singlets remain strongly deshielded in the range of 10.08−10.51 δ. In 1H NMR spectra of nucleosides 22A(a−g), the signals corresponding to the anthracene and isoxazolidine moieties are



DISCUSSION AND CONCLUSIONS We have investigated the ene reactions of aromatic nitrosocarbonyls, generated through the mild oxidative protocol with NMO, with allylic alkoxy olefins. The faster oxidation process of nitrile oxide to nitrosocarbonyl intermediate prevents a possible side reaction, that is, the addition of the allylic alcohol to the electrophilic nitrile oxide. From the reaction mixtures, the presence of adducts between the aromatic nitrile oxides and the 3-methyl-2-buten-1-ol 9d was not observed. The nitrosocarbonyl benzene 8P adds the 3-methyl-2-buten1-ol 9d, affording adducts 10Pd and 13, which derive from the preferred M path in ene reactions of trisubstituted olefins.13,16 If hydroxy-protected allylic ether 9e or ester 9f are used, the reactions proceed straightforwardly to single ene adducts 10Pe,f in accordance with the prevailing HOMO(olefin)− LUMO(nitrosocarbonyl) interaction, somewhat enforced by the polarization of the CC double bond induced by the CH2OR group.36 Furthermore, no hydrogen-bonding effects are found in directing the selectivity in the reactions at hand. When bulkier substituents replace the phenyl group (Figure 1), as in nitrosocarbonyl intermediates 8M and 18, in the ene reactions with the 3-methyl-2-buten-1-ol 9d, the M path is risen energetically because of a steric effect and the AM path is favored, affording a mixture of regioisomeric compounds. We have recently detailed the selectivity outcome in ene reactions of nitrosocarbonyl mesitylene 8M with trisubstituted olefins,16 and the results show the remarkable influence of the nitrosocarbonyl substituent on the selectivities in ene reactions. Model calculations on the reaction of 8P,M with tetramethylethylene (TME) shed light on the factors involved in the 520

dx.doi.org/10.1021/jo302346a | J. Org. Chem. 2013, 78, 516−526

The Journal of Organic Chemistry

Article

group is twisted out of the nitrosocarbonyl plane as usual, causing unfavorable steric crowding between its ortho methyl and the “trans” distal TME methyl, as depicted in Figure 2. This crowding should slow down the M approach of 8M to a trisubstituted ethene and compensates somewhat the M electronic bias. Figure 3 reports the same features in the case of the allylic alcohol 9d. From the M approach point of view, it seems quite clear that the OH group is not playing any role in directing the addition of the nitrosocarbonyl intermediates, both in the cases depicted in the TS M and M′, as the top view of the structures shows (Figure 3). The steric clashes at work in the M case drift the addition to the AM path that is the prelude to the enol formation and subsequent cyclization to the isoxazolidine structures. The structures of the 5-hydroxy-isozaxolidine 11 and 20 resemble those obtainable through nitrone additions to vinyl ether derivatives and constitute useful synthons toward the prepartion of N,O-nucleoside analogues. By adapting the Vorbrüggen protocol23 as well as known procedures for similar compounds,5,24 we have prepared a selection of uracil derivatives 22M,A(a−g) from good to excellent yields. These products constitute a new class of nucleoside analogues with a variety of different heterobases, easily inserted on the isoxazolidine ring. Their potentialities in terms of biological activity are totally unexplored, but we believe in future promising results if we compare the new nucleoside structures with those of known compounds, both hetero- and carbocyclic, recently reported in the literature, which were found to be interesting adenosine receptor agonists.37 The insertion of a fluorecent chromophore allows for planning future applications as fluorescent markers in different biological targets. The uracil derivatives 22M,A(a−g) were sent to the Southern Research Institute (SRI) of Birmingham (AL, USA) within the Tuberculosis Antimicrobial Acquisition and Coordinating Facility (TAACF) program and tested against the Mycobacterium tuberculosis H37Rv in BACTEC 12B medium using the microplate Alamar Blue Assay.38 The minimum inhibition concentration (MIC μg/mL) was found to be >6.25 for all the compounds with the %Inh in the range of 22−34. Further antiviral and antitumoral tests are currently under evaluation.

Figure 1. Effect of substituents on ene reaction selectivity.

varying selectivities. In these reactions of trimethylethylenes, the M path is favored in the case of nitrosocarbonyl benzene 8P while steric hindrance in the approach of 8M compensates somewhat its electronic preference, and mixtures of M and AM adducts are formed. This mechanism is at work also in the case of 3-methyl-2-buten-1-ol 9d as well as the corresponding hydroxy-protected derivatives 9e,f (Figure 2) and remarkably influences and enforces the selectivity in the reaction of nitrosocarbonyl anthracene 18 with 3-methyl-2-buten-1-ol 9d. The steric hindrance produced by the anthracene group reverses the selectivity in the ene reaction, and the corresponding 5-hydroxy-isoxazolidine becomes the major product and a convenient intermediate for nucleoside analogue synthesis. As we have demonstrated,16 in the reactions of nitrosocarbonyl 8P with TME, the transition structure (TS) of the addition step of 8P to TME shows no special hindrance between the addends, whereas in the TS of 8M, the mesityl

Figure 2. Side view of the B3LYP/6-31G* TS's for the reactions of tetramethylethylene with nitrosocarbonyl benzene and 2,6-dimethylbenzene.16 521

dx.doi.org/10.1021/jo302346a | J. Org. Chem. 2013, 78, 516−526

The Journal of Organic Chemistry

Article

Figure 3. Side views and top views of the M and AM ene additions to the 3-methyl-2-buten-1-ol of the nitrosocarbonyl mesitylene.



10Pd: (1.42 g, 50%) White crystals from ethanol, mp 157−159 °C. IR: νOH 3450, νCO 1715 cm−1. 1H NMR (DMSO) δ: 1.70 (s, 3H, CH3); 4.70 (m, 2H, CH2-O); 5.15 (s, 1H + 1H, CH2); 5.30 (m, 1H, CH-N); 7.2−7.9 (m, 5H, Ph); 8.21 (s, 1H, OH). 13C NMR (DMSO) δ: 23.8; 66.0; 79.9; 120.4; 132.6; 132.9; 133.1; 133.2; 137.5; 142.7; 169.0. Anal. Calcd for C12H15NO3 (221.25): C, 65.14; H, 6.83; N, 6.33. Found: C, 65.15; H, 6.83; N, 6.25. 10Pe: (2.12 g, 70%) Yellowish oil. IR: νOH 3203, νCO 1626 cm−1. 1 H NMR (DMSO) δ: 1.75 (s, 3H, CH3); 3.33 (s, 3H, OCH3); 3.56 (dd, J = 10, 4 Hz, 2H, CH-N); 3.74 (t, J = 10 Hz, 2H, CH2-O); 4.89 and 4.97 (s, 1H + 1H, CH2); 7.3−7.6 (m, 5H, Ph); 9.50 (s, 1H, OH). 13 C NMR (DMSO) δ: 21.0; 58.1; 69.2; 113.4; 126.0; 127.8; 128.0; 128.2; 128.5; 129.2; 135.4; 140.9; 169.2. Anal. Calcd for C13H17NO3 (235.27): C, 66.36; H, 7.28; N, 5.95. Found: C, 66.35; H, 7.25; N, 5.94. 10Pf: (1.81 g, 60%) Yellowish oil. IR: νOH 3206, νCO 1742, 1612 cm−1. 1H NMR (DMSO at 80 °C) δ: 1.80 (s, 3H, CH3); 2.01 (s, 3H, CH3CO); 4.35 (d, J = 7 Hz, 2H, CH-O); 4.91 and 5.00 (s, 1H + 1H, CH2); 5.03 (m, 1H, CH-N); 7.39 (m, 3H, arom.); 7.57 (m, 2H, arom.); 9.35 (s, 1H, OH). 13C NMR (DMSO at 80 °C) δ: 13.5; 20.4; 57.8; 61.9; 113.5; 126.0; 128.3; 128.7; 129.1; 130.0; 140.6; 165.5; 171.5. Anal. Calcd for C13H17NO3 (235.27): C, 66.36; H, 7.28; N, 5.95. Found: C, 66.34; H, 7.25; N, 5.94. 13Pd: (0.21 g, 8%) Yellowish oil. IR: νOH 3393, νCO 1723 cm−1. 1 H NMR (CDCl3) δ: 1.90 (s, 3H, CH3); 5.22 (m, 2H, CH2); 5.32 and 5.43 (s, 1H + 1H, CH2); 7.2−8.0 (m, 5H, Ph). 13C NMR (CDCl3) δ: 19.3; 54.1; 118.6; 128.3; 129.3; 129.6; 133.0; 138.5; 154.9; 166.1. Anal. Calcd for C12H13NO2 (203.23): C, 70.91; H, 6.45; N, 6.89. Found: C, 70.85; H, 6.45; N, 6.85. General Procedure for the Ene Reactions of Nitrosocarbonyl Mesitylene 8M with Alkenes 9d−9f. To an ice-cooled DCM (200 mL) solution of alkenes 9d−9f (5 equiv) was added 1.8 g (1.3 equiv) of NMO under stirring. A solution of 2.0 g (12.4 mmol) of mesitonitrile oxide 7 in 100 mL of DCM was added dropwise, and the reaction was left under stirring at r.t. for 24 h. After dilution with an equivalent volume of DCM, the organic phase was washed with water and dried over anhydrous Na2SO4. After filtration, the solvent was

EXPERIMENTAL SECTION

All melting points (mp's) are uncorrected. Elemental analyses were done on n elemental analyzer available at the Departmen of Chemistryt. 1H and 13C NMR spectra were recorded on a 300 MHz spectrometer (solvents specified). Chemical shifts are expressed in parts per million (ppm) from internal tetramethylsilane (δ), and coupling constants (J) are in hertz (Hz): b, broad; s, singlet; bs, broad singlet; d, doublet; t, triplet; m, multiplet. IR spectra (Nujol mulls) were recorded on a spectrophotometer available at the Department, and absorbtions (ν) are in cm−1. Column chromatography and tlc: silica gel H60 and GF254, respectively. Eluants: cyclohexane/ethyl acetate 9:1 to pure ethyl acetate; pure CHCl3 to CHCl3/MeOH 9/1 for the nucleoside syntheses. The identification of samples from different experiments was secured by mixed mp's and superimposable IR spectra. Starting and Reference Materials. 3-Methyl-2-buten-1-ol (99%) 9d was purchased from chemical suppliers. 3-Methyl-2-butenyl methyl ether 9e was prepared from 4-bromo-2methyl-2-butene and MeONa according to the reported synthesis.17 1-Acetoxy-3-methyl-2-butene 9f was prepared from the alcohol 9d and acetic anhydride as reported.18 Benzhydroximoyl chloride was obtained by treatment of benzaldoxime with sodium hypochlorite.39 Addition of a slight excess of Et3N to a DCM solution of benzhydroximoyl chloride furnished in situ BNO. Mesitonitrile oxide 7 was obtained by oxidation of 2,4,6trimethylbenzaldoxime with bromine.21 The 9-anthraldeyde oxime 16 has been purchased from chemical suppliers. General Procedure for the Ene Reactions of Nitrosocarbonyl Benzene 8A with Alkenes 9d−9f. To an ice-cooled DCM (200 mL) solution of alkenes 9d−9f (5 equiv) were added 1.9 g (1.3 equiv) of NMO and 2 mL (1.1 equiv) of Et3N under stirring. A solution of 2.0 g (12.9 mmol) of benzhydroxymoyl chloride 6 in 100 mL of DCM was added dropwise, and the reaction was left under stirring at r.t. for 24 h. After dilution with an equivalent volume of DCM, the organic phase was washed with water and dried over anhydrous Na2SO4. After filtration, the solvent was then evaporated and the reaction mixture was separated by column chromatography, affording the ene adducts. 522

dx.doi.org/10.1021/jo302346a | J. Org. Chem. 2013, 78, 516−526

The Journal of Organic Chemistry

Article

20: (1.52 g, 52%) Straw yellow crystals from ethanol, mp 229 °C (dec.). IR: νOH 3200, νCO 1588 cm−1. 1H NMR (DMSO) δ: 1.83 (s, 3H, CH3); 1.97 (s, 3H, CH3); 2.22 (d, 1H, J = 13 Hz, CH); 2.44 (dd, 1H, J = 13, 5 Hz, CH); 5.23 (b, 1H, CH-OH); 6.84 (b, 1H, OH, exch. with D2O); 7.56 (m, 4H, arom.); 7.89 (m, 1H, arom.); 7.99 (m, 1H, arom.); 8.11 (m, 2H, arom.); 8.62 (s, 1H, arom.). 13C NMR (DMSO) δ: 25.7; 26.6; 50.0; 61.6; 95.8; 124.8; 125.3; 125.5; 125.6; 126.3; 126.4; 126.6; 126.7; 126.8; 128.2; 128.4; 130.6; 130.8; 131.9; 161.8. Anal. Calcd for C20H19NO3 (321.38): C, 74.75; H, 5.96; N, 4.36. Found: C, 74.76; H, 6.00; N, 4.37. Synthesis of the N-Mesitoyl-3,3-dimethyl-5-acetoxy-1,2isoxazolidine 21M. To an ice-cooled anhydrous DCM (50 mL) solution of N-mesitoyl-3,3-dimethyl-5-hydroxy-1,2-isoxazolidine 11Md (3.60 g, 14 mmol.) was added 2.2 equiv of Ac2O added under stirring along with 0.34 equiv of DMAP and 2.2 equiv of Et3N. The reaction was left under stirring at r.t. for 24 h. After dilution with an equivalent volume of DCM, the organic phase was washed with a saturated solution of NaHCO3 and dried over anhydrous Na2SO4. After filtration, the solvent was evaporated and a solid was obtained. The acetyl derivative 21M (3.68 g) was obtained in 86% yield and recrystallized from methanol: mp 105−107 °C. IR: νCO 1760 cm−1. 1 H NMR (CDCl3) δ: 1.74 and 1.88 (s, 6H, 2CH3); 2.09 (s, 3H, CH3CO); 2.28 (s, 9H, 3CH3-Ar); 2.50 (m, 2H, CH2); 6.23 (d, J = 6 Hz, 1H, O-CH-OAc); 6.80 (m, 2H, Ph). 13C NMR (CDCl3) δ: 18.4; 18.6; 20.9; 21.0; 49.8; 60.6; 93.9; 127.7; 127.8; 132.8; 133.5; 138.0; 165.1; 169.1. Anal. Calcd for C17H23NO4 (305.36): C, 66.86; H, 7.59; N, 4.59. Found: C, 66.75; H, 7.56; N, 4.58. Synthesis of the N-Anthracenoyl-3,3-dimethyl-5-acetoxy1,2-isoxazolidine 21A. To an ice-cooled anhydrous DCM (50 mL) solution of N-anthracenoyl-3,3-dimethyl-5-hydroxy-1,2-isoxazolidine 20 (1.40 g, 4.78 mmol.) was added 2.2 equiv of Ac2O under stirring along with 0.34 equiv of DMAP and 2.2 equiv of Et3N. The reaction was left under stirring at r.t. for 24 h. After dilution with an equivalent volume of DCM, the organic phase was washed with a saturated solution of NaHCO3 and dried over anhydrous Na2SO4. The solvent was then evaporated, and a solid was obtained. The acetyl derivative 21A (1.63 g) was obtained in 94% yield and recrystallized from ethanol: mp 200−201 °C. IR: νCO 1748 cm−1. 1H NMR (CDCl3) δ: 1.84 and 1.97 (s, 3H + 3H, CH3); 2.15 (s, 3H, CH3CO); 2.47 (d, 1H, J = 13 Hz, H-CH); 2.64 (dd, 1H, J = 13, 5 Hz, H-CH); 6.04 (d, 1H, J = 5 Hz, O-CH-O); 7.51 (m, 4H, arom.); 8.04 (m, 4H, arom.); 8.47 (m, 1H, arom.). 13C NMR (CDCl3) δ: 20.8; 26.0; 27.2; 49.5; 61.5; 93.5; 124.7; 124.9; 125.3; 125.8; 126.4; 127.0; 127.5; 127.8; 128.5; 128.6; 130.7; 130.9; 131.1; 164.0; 168.8. Anal. Calcd for C22H21NO4 (363.40): C, 72.71; H, 5.82; N, 3.85. Found: C, 72.71; H, 5.81; N, 3.89. General Procedure for the Synthesis of Nucleosides 22 by Coupling of Isoxazolidines 21M,A and Heterobases B(a−g). A solution of 2 equiv of heterobases B and 2 equiv of bis(trimethylsilyl)acetamide (BSA) in anhydrous DCM (50 mL) is refluxed under a nitrogen atmosphere for 15−20 min until it becomes clear and hence cooled to ambient. A solution in DCM (10 mL) of isoxazolidines 21M,A (0.20 g, 0.65 mmol) is added dropwise and cooled to 0 °C and additioned of 0.12 mL (1 equiv) of TMSO-Tf. The reaction is refluxed under stirring overnight and finally quenched with a saturated solution of NaHCO3 at pH = 7. The mixture is diluted with an equivalent volume of DCM and washed with water and finally dried over Na2SO4. From the residues, nucleosides 22M,A(a−g) are isolated upon crystallization or column chromatographic purification. 22Ma: (0.22 g, 95%) White solid from ethanol, mp 215 °C (dec.). IR: νNH 3166, νCO 1690, 1642 cm−1. 1H NMR (CD3COCD3) δ: 1.75 and 1.89 (s, 6H, 2CH3); 2.22 and 2.25 (s, 9H, 3CH3-Ar); 2.72 and 2.90 (dd, J = 14, 7 Hz, 2H, CH2); 5.67 (d, J = 8 Hz, 1H, CO-CH); 6.40 (t, J = 7 Hz, 1H, CH); 6.82 (m, 2H, Ph); 7.64 (d, J = 8, 1H, CHN); 10.10 (bs, 1H, NH). 13C NMR (CD3COCD3) δ: 19.5; 21.4; 25.4; 26.9; 48.8; 64.3; 83.4; 103.9; 129.0; 134.4; 134.9; 139.1; 140.5; 151.6; 154.5; 163.4; 168.0; 170.0. Anal. Calcd for C19H23N3O4 (357.40): C, 63.85; H, 6.48; N, 11.76. Found: C, 63.83; H, 6.45; N, 11.75. 22Mb: (0.24 g, 97%) White crystals from acetone, mp 210 °C (dec.). IR: νNH 3167, νCO 1709 cm−1. 1H NMR (CD3COCD3) δ:

then evaporated and the reaction mixture was separated by column chromatography, affording the ene adducts. 10Md: (1.63 g, 50%) White crystals from ethanol, mp 162−163 °C. IR: νOH 3163, νCO 1618 cm−1. 1H NMR (DMSO) δ: 1.78 (s, 3H, CH3); 2.14, 2.18, and 2.21 (s, 9H, 3CH3-Ar); 3.72 (dd, J = 7, 6 Hz, 2H, CH2-O); 4.68 (m, 1H, CH-N); 4.97 (bs, 1H + 1H, CH2); 5.06 (t, J = 7 Hz, 1H, OH); 6.80 (m, 2H, Ph); 9.04 (s, 1H, OH). 13C NMR (DMSO) δ: 18.8; 18.9; 20.6; 21.4; 58.8; 60.5; 113.4; 127.3; 133.3; 133.5; 134.6; 136.5; 141.2; 170.0. Anal. Calcd for C15H21NO3 (263.33): C, 68.41; H, 8.04; N, 5.32. Found: C, 68.47; H, 8.05; N, 5.25. 11Md: (0.82 g, 25%) White crystals from benzene, mp 157−158 °C. IR: νOH 3150, νCO 1612 cm−1. 1H NMR (DMSO) δ: 1.55 and 1.72 (s, 6H, 2CH3); 2.13, 2.19, and 2.22 (s, 9H, 3CH3−Ar); 2.14 and 2.35 (s, 1H + 1H, CH2); 5.44 (m, 1H, CH); 6.79 (m, 2H, Ph); 6.98 (s, 1H, OH). 13C NMR (DMSO) δ: 18.3; 20.6; 25.5; 26.4; 50.2; 60.6; 95.3; 127.2; 127.5; 132.8; 132.9; 134.9; 136.7; 163.2. Anal. Calcd for C15H21NO3 (263.33): C, 68.41; H, 8.04; N, 5.32. Found: C, 68.47; H, 8.06; N, 5.26. 10Me: (2.17 g, 63%) White crystals from ethanol, mp 127−128 °C. IR: νOH 2920, νCO 1603 cm−1. 1H NMR (DMSO) δ: 1.79 (s, 3H, CH3); 2.16, 2.17, and 2.21 (s, 9H, 3CH3-Ar); 3.29 (s, 3H, CH3-O); 3.70 (m, 2H, CH2-O); 4.90 and 5.00 (s, 1H + 1H, CH2); 5.24 (m, 1H, CH-N); 6.81 (m, 2H, Ph); 9.07 (s, 1H, OH). 13C NMR (DMSO) δ: 18.6; 18.8; 20.1; 20.6; 21.3; 56.8; 57.8; 68.9; 113.4; 127.3; 133.3; 133.5; 134.5; 136.5; 140.7; 170.0. Anal. Calcd for C16H23NO3 (277.35): C, 69.28; H, 8.36; N, 5.05. Found: C, 69.26; H, 8.25; N, 5.08. 12Me: (1.20 g, 35%) White crystals from ethanol, mp 133−135 °C. IR: νOH 2920, νCO 1656 cm−1. 1H NMR (DMSO) δ: 1.51 (s, 6H, CH3); 2.12 and 2.20 (s, 9H, 3CH3-Ar); 3.46 (s, 3H, CH3-O); 5.23 (d, J = 13 Hz, 1H, C-CH); 6.52 (d, J = 13 Hz, 1H, O-CH); 6.80 (m, 2H, Ph); 9.10 (s, 1H, OH). 13C NMR (DMSO) δ: 18.5; 20.6; 26.6; 55.5; 61.0; 108.6; 127.2; 132.5; 133.7; 135.9; 136.4; 146.8; 154.5; 170.8. Anal. Calcd for C16H23NO3 (277.35): C, 69.28; H, 8.36; N, 5.05. Found: C, 69.25; H, 8.37; N, 5.07. 10Mf: (2.12 g, 56%) White crystals from ethanol, mp 103−104 °C. IR: νOH 3086, νCO 1747, 1658 cm−1. 1H NMR (DMSO) δ: 1.80 (s, 3H, CH3); 2.00 (s, 3H, CH3CO); 2.18 and 2.22 (s, 9H, 3CH3-Ar); 4.25 and 4.41 (AB syst, 2H, CH2-O); 4.96 and 5.04 (s, 1H + 1H, CH2); 5.30 (dd, J = 10, 4 Hz, 1H, CH-N); 6.82 (m, 2H, Ph); 9.27 (s, 1H, OH). 13C NMR (DMSO) δ: 18.7; 18.8; 20.6; 20.7; 21.2; 56.4; 60.8; 114.0; 127.3; 127.4; 133.3; 133.4; 134.2; 136.7; 140.0; 170.1; 170.2. Anal. Calcd for C17H23NO4 (305.36): C, 66.86; H, 7.59; N, 4.59. Found: C, 66.87; H, 7.58; N, 5.00. 12Mf: (1.21 g, 32%) White crystals from ethanol, mp 108−109 °C. IR: νOH 3098, νCO 1754, 1667 cm−1. 1H NMR (DMSO) δ: 1.52 (s, 6H, CH3); 2.12 and 2.20 (s, 9H, 3CH3-Ar); 2.12 (s, 3H, CH3CO); 5.88 (d, J = 13 Hz, 1H, C-CH); 6.79 (m, 2H, Ph); 7.21 (d, J = 13 Hz, 1H, O-CH); 9.38 (s, 1H, OH). 13C NMR (DMSO) δ: 18.5; 20.5; 20.6; 25.8; 60.4; 120.3; 127.2; 132.7; 134.5; 135.9; 136.2; 168.0; 170.8. Anal. Calcd for C17H23NO4 (305.36): C, 66.86; H, 7.59; N, 4.59. Found: C, 66.85; H, 7.55; N, 5.00. Ene Reaction of Anthracenenitrile Oxide 18 with 3-Methyl2-Buten-1-ol 9d. To an anhydrous DCM (80 mL) solution of 3methyl-2-buten-1-ol 9d (5 equiv) and NMO (2 equiv) was added 2.0 g (9 mmol) of anthracenenitrile oxide 18 portionwise under stirring at r.t. After 48 h, the organic phase was washed with brine and dried over anhydrous Na2SO4. After filtration and evaporation of the solvent, the residue was submitted to column chromatography. The two main fractions were identified as the adducts 19 and 20, which were characterized as below: 19: (0.79 g, 27%) Pale yellow crystals from ethanol, mp 179−181 °C. IR: νOH 3254, νCO 1603 cm−1. 1H NMR (DMSO) δ: 1.80 (s, 3H, CH3); 3.35 (s, 2H, CH2OH); 4.41 (m, 3H, OH); 4.67 (m, 2H, CH2); 5.49 (b, 1H, CH); 7.57 (m, 4H, arom.); 8.11 (m, 4H, arom.); 8.59 (s, 1H, arom.); 9.71 (s, 1H, OH). 13C NMR (DMSO) δ: 17.9; 25.5; 61.1; 119.6; 123.6; 125.5; 125.7; 126.0; 126.7; 128.4; 129.1; 131.3; 134.6; 155.6. Anal. Calcd for C20H19NO3 (321.38): C, 74.75; H, 5.96; N, 4.36. Found: C, 74.69; H, 5.94; N, 4.36. 523

dx.doi.org/10.1021/jo302346a | J. Org. Chem. 2013, 78, 516−526

The Journal of Organic Chemistry

Article

45.9; 63.8; 82.7; 124.0; 124.3; 124.4; 124.6; 125.7; 126.4; 126.6; 126.8; 126.9; 127.8; 128.5; 129.6; 130.5; 130.6; 138.2; 138.4; 141.5; 148.8; 156.5; 156.8; 164.2. Anal. Calcd for C24H20O4N3F (433.42) C, 66.51; H, 4.65; N, 9.70. Found: C, 66.51; H, 4.64; N, 9.71. 22Ac: (0.29 g, 97%) White solid from ethanol, mp 185−188 °C. IR: νNH 3067, νCO 1728 cm−1. 1H NMR (DMSO) δ: 1.83 (s, 3H, CH3); 2.03 (s, 3H, CH3); 2.84 (m, 2H, CH2); 6.12 (1H, t, J = 6 Hz, O-CHN); 7.58 (m, 4H, arom.); 7.72 (s, 1H, N-CH); 7.86 (d, 1H, J = 8 Hz, arom.); 8.00 (d, 1H, J = 8 Hz, arom.); 8.12 (m, 2H, arom.); 8.67 (s, 1H, arom.); 11.78 (s, 1H, NH). 13C NMR (DMSO) δ: 24.8; 25.7; 46.3; 63.7; 83.1; 107.9; 124.3; 125.6; 125.7; 126.4; 126.8; 126.9; 127.9; 128.6; 130.4; 130.5; 130.6; 136.7; 149.2; 158.6; 164.4. Anal. Calcd for C24H20O4N3Cl (449.89) C, 64.07; H, 4.48; N, 9.34. Found: C, 64.08; H, 4.49; N, 9.31. 22Ad: (0.31 g, 95%) White solid from ethanol, mp 167−170 °C. IR: νNH 3058, νCO 1726 cm−1. 1H NMR (DMSO) δ: 1.83 (s, 3H, CH3); 2.03 (s, 3H, CH3); 2.85 (m, 2H, CH2); 6.12 (t, 1H, J = 6 Hz, O-CH-N); 7.57 (m, 4H, arom.); 7.81 (s, 1H, N-CH); 7.86 (d, 1H, J = 8 Hz, arom.); 7.97 (d, 1H, J = 8, 4 Hz, arom.); 8.13 (m, 2H, arom.); 8.67 (s, 1H, arom.); 11.73 (s, 1H, NH). 13C NMR (DMSO) δ: 24.9; 25.7; 46.1; 63.7; 83.1; 96.5; 124.3; 125.6; 126.4; 126.9; 127.8; 128.6; 130.4; 130.5; 139.2; 149.4; 158.7; 164.4. Anal. Calcd for C24H20O4N3Br (494.33) C, 58.31; H, 4.08; N, 8.50. Found: C, 58.31; H, 4.11; N, 8.51. 22Ae: (0.35 g, 99%) White solid from ethanol, mp 164−166 °C. IR: νNH 3059, νCO 1724 cm−1. 1H NMR (DMSO) δ: 1.83 (s, 3H, CH3); 2.03 (s, 3H, CH3); 2.86 (m, 2H, CH2); 6.13 (t, 1H, J = 6 Hz, O-CHN); 7.58 (m, 4H, arom.); 7.85 (d, 2H, J = 7 Hz, arom.); 7.86 (s, 1H, N-CH); 7.96 (d, 1H, J = 8 Hz, arom.); 8.13 (m, 2H, arom.); 8.67 (s, 1H, arom.); 11.60 (s, 1H, NH). 13C NMR (DMSO) δ: 25.0; 25.7; 45.9; 63.7; 70.6; 82.8; 124.3; 125.6; 126.4; 126.9; 127.0; 127.8; 128.6; 130.4; 130.5; 143.9; 149.8; 160.1; 164.4. Anal. Calcd for C24H20O4N3I (541.32) C, 53.25; H, 3.72; N, 7.76. Found: C, 53.22; H, 3.72; N, 7.80. 22Af: (0.24 g, 86%) White solid from ethanol, mp 155−157 °C. IR: νNH 3058, νCO 1719 cm−1. 1H NMR (DMSO) δ: 1.63 (s, 3H, CH3); 1.84 (s, 3H, CH3); 2.05 (s, 3H, CH3); 2.78 (m, 2H, CH2); 6.19 (t, 1H, J = 7 Hz, O-CH-N); 7.29 (d, 1H, J = 1 Hz, N-CH); 7.59 (m, 4H, arom.); 7.86 (d, 1H, J = 9 Hz); 8.04 (m, 1H, arom.); 8.13 (m, 2H, arom.); 8.66 (s, 1H, arom.); 11.23 (s, 1H, NH). 13C NMR (DMSO) δ: 11.5; 24.4; 25.5; 45.9; 63.3; 81.6; 109.7; 123.9; 124.3; 125.2; 126.0; 126.2; 126.4; 126.5; 127.4; 128.1; 128.2; 130.0; 130.1; 134.4; 149.7; 162.9; 163.8. Anal. Calcd for C25H23O4N3 (429.46) C, 69.91; H, 5.40; N, 9.79. Found: C, 69.91; H, 5.41; N, 9.75. 22Ag: (0.04 g, 48%) White solid from ethanol, mp 188−191 °C. IR: νNH2 3362, 3171, νCO 1664 cm−1. 1H NMR (DMSO) δ: 1.87 (s, 3H, CH3); 1.98 (s, 1.98, CH3); 2.59 e 2.84 (m, 1H + 1H, CH2); 5.53 (d, 1H, J = 7 Hz, CH-CN); 6.22 (t, 1H, J = 6 Hz, O-CH-N); 7.18 (d, 2H, NH2); 7.39 (d, 1H, J = 7 Hz, N-CH); 7.58 (m, 4H, arom.); 7.88 (d, 1H, J = 8 Hz, arom.); 8.00 (m, 1H, arom.); 8.13 (m, 2H, arom.); 8.66 (s, 1H, arom). 13C NMR (DMSO) δ: 22.8; 25.1; 26.0; 47.2; 63.4; 82.7; 94.7; 124.3; 124.8; 125.6; 125.7; 126.4; 126.6; 126.8; 126.9; 127.7; 128.5; 128.6; 130.5; 130.6; 140.0; 154.4; 163.8; 165.3. Anal. Calcd for C24H22O3N4 (414.45) C, 69.55; H, 5.35; N, 13.52. Found: C, 69.51; H, 5.31; N, 13.52. Competition Experiments. Equimolecular amounts (20 equiv) of 3-methyl-2-buten-1-ol and the corresponding acetate and methyl ether were allowed to react with 1 equiv of either benzonitrile oxide or mesitonitrile oxide in the presence of 1.2 equiv of NMO in anhydrous DCM as solvent. After the workup previously described, the reaction mixtures were submitted to chromatographic separation for the isolation and quantification of the ene obtained from the yields given in Table 1 (see Table 2). Biological Tests. Primary screen (dose response): determination of a 90% inhibitory concentration (IC90). The initial screen is conducted against Mycobacterium tuberculosis H37Rv (ATCC 27294) in BACTEC 12B medium using the microplate Alamar Blue assay (MABA). Compounds are tested in ten 2-fold dilutions, typically from 100 to 0.19 μg/mL. The IC90 is defined as the concentration effecting a reduction in fluorescence of 90% relative to controls. This value is determined from the dose−response curve using a curve-fitting

1.74 and 1.88 (s, 6H, 2CH3); 2.24, 2.25, and 2.28 (s, 9H, 3CH3-Ar); 2.76 and 2.90 (dd, J = 14, 7 Hz, 2H, CH2); 6.35 (dt, J = 7, 1 Hz, 1H, CH); 6.84 (m, 2H, Ph); 7.70 (d, J = 7 Hz, 1H, CH-N); 10.50 (bs, 1H, NH). 13C NMR (CD3COCD3) δ: 19.4; 19.5; 21.5; 25.3; 26.9; 49.0; 64.3; 84.1; 124.3; 124.8; 129.1; 134.4; 135.0; 135.4; 139.2; 140.5; 143.6; 150.2; 157.5; 157.8; 167.9. Anal. Calcd for C19H22N3O4F (375.40): C, 60.79; H, 5.91; N, 11.19. Found: C, 60.81; H, 5.91; N, 11.21. 22Mc: (0.24 g, 93%) White crystals from acetone, mp 208−211 °C. IR: νNH 3167, νCO 1709 cm−1. 1H NMR (CD3COCD3) δ: 1.73 and 1.87 (s, 6H, 2CH3); 2.25, 2.26, and 2.31 (s, 9H, 3CH3-Ar); 2.83 and 2.95 (dd, J = 14, 7 Hz, 2H, CH2); 6.29 (dt, J = 7, 1 Hz, 1H, CH); 6.85 and 6.89 (1, 1H + 1H, Ph); 7.68 (s, 1H, CH-N); 10.60 (bs, 1H, NH). 13 C NMR (CD3COCD3) δ: 19.4; 19.5; 21.5; 25.1; 49.7; 64.2; 84.5; 109.8; 129.1; 129.2; 134.2; 135.0; 135.3; 137.3; 139.3; 150.6; 159.6; 168.2. Anal. Calcd for C19H22N3O4Cl (391.85): C, 58.24; H, 5.66; N, 10.72. Found: C, 58.22; H, 5.69; N, 10.71. 22Md: (0.27 g, 95%) White crystals from methanol, mp 114−120 °C. IR: νNH 3168, νCO 1700 cm−1. 1H NMR (CD3COCD3) δ: 1.72 and 1.86 (s, 6H, 2CH3); 2.25 and 2.32 (s, 9H, 3CH3-Ar); 2.84 and 2.96 (dd, J = 14, 7 Hz, 2H, CH2); 6.28 (t, J = 6 Hz, 1H, CH); 6.87 (m, 2H, Ph); 7.74 (s, 1H, CH-N); 10.51 (bs, 1H, NH). 13C NMR (CD3COCD3) δ: 19.5; 19.6; 21.5; 25.1; 26.9; 49.8; 64.2; 84.6; 97.8; 129.0; 129.3; 134.2; 135.0; 135.3; 139.3; 139.8; 150.8; 159.6; 168.3. Anal. Calcd for C19H22N3O4Br (436.30): C, 52.31; H, 5.08; N, 9.63. Found: C, 52.25; H, 5.09; N, 9.55. 22Me: (0.29 g, 91%) White crystals from methanol, mp 106−110 °C. IR: νNH 3169, νCO 1690 cm−1. 1H NMR (CD3COCD3) δ: 1.72 and 1.87 (s, 6H, 2CH3); 2.25 and 2.34 (s, 9H, 3CH3-Ar); 2.82 and 2.96 (dd, J = 14, 7 Hz, 2H, CH2); 6.27 (t, J = 6 Hz, 1H, CH); 6.85 (m, 2H, Ph); 7.80 (s, 1H, CH-N); 10.45 (bs, 1H, NH). 13C NMR (CD3COCD3) δ: 19.5; 19.8; 21.5; 25.0; 26.9; 49.9; 64.2; 69.9; 84.4; 129.0; 129.4; 134.1; 135.0; 135.3; 139.2; 144.7; 151.2; 160.8; 168.4. Anal. Calcd for C19H22N3O4I (483.30): C, 47.22; H, 4.59; N, 8.69. Found: C, 47.15; H, 4.55; N, 8.49. 22Mf: (0.19 g, 78%) White crystals from ethyl acetate, mp 99−104 °C. IR: νNH 3170, νCO 1700 cm−1. 1H NMR (CD3COCD3) δ: 1.74 and 1.88 (s, 6H, 2CH3); 1.83 (s, 3H, CH3-C); 2.22, 2.25, and 2.27 (s, 9H, 3CH3-Ar); 2.74 and 2.86 (dd, J = 14, 7 Hz, 2H, CH2); 6.39 (t, J = 7 Hz, 1H, CH); 6.82 (m, 2H, Ph); 7.43 (d, J = 1 Hz, 1H, CH-N); 10.08 (bs, 1H, NH). 13C NMR (CD3COCD3) δ: 12.8; 19.5; 21.4; 25.3; 26.9; 48.6; 64.3; 83.3; 112.1; 129.0; 134.3; 134.9; 135.5; 135.8; 139.0; 151.6; 164.2; 167.7. Anal. Calcd for C20H25N3O4 (371.42): C, 64.67; H, 6.78; N, 11.31. Found: C, 64.65; H, 6.75; N, 11.25. 22Mg: (0.11 g, 46%) White crystals from ethyl acetate, mp 107− 110 °C. IR: νNH2 3360, 3170, νCO 1640 cm−1. 1H NMR (CD3COCD3) δ: 1.75 and 1.83 (s, 6H, 2CH3); 2.21, 2.24, and 2.26 (s, 9H, 3CH3-Ar); 2.60 and 2.83 (dd, J = 14, 7 Hz, 2H, CH2); 5.88 (d, J = 8 Hz, 1H, CH); 6.43 (t, J = 7 Hz, 1H, CH); 6.81 (m, 2H, Ph); 7.38 (s, 2H, NH2); 7.56 (d, J = 8 Hz, 1H, CH-N). 13C NMR (CD3COCD3) δ: 19.5; 21.4; 23.6; 25.6; 27.1; 49.5; 64.0; 84.2; 95.7; 128.8; 128.9; 129.0; 129.5; 134.3; 134.9; 135.5; 139.0; 141.4; 156.1; 167.3. Anal. Calcd for C19H24N4O3 (356.41): C, 64.02; H, 6.79; N, 15.72. Found: C, 63.90; H, 6.65; N, 15.55. 22Aa: (0.26 g, 97%) White solid from ethanol, mp > 200 °C. IR: νNH 3167, νCO 1686 cm−1. 1H NMR (DMSO) δ: 1.84 (s, 3H, CH3); 2.03 (s, 3H, CH3); 2.79 (m, 2H, CH2); 5.49 (d, 1H, J = 8 Hz, CHCO); 6.17 (dd, 1H, J = 7, 6 Hz, O-CH-N); 7.59 (m, 5H, arom.); 7.86 (d, 1H, J = 8 Hz, N-CH); 7.99 (m, 1H, arom.); 8.12 (m, 2H, arom.); 8.66 (s, 1H, arom.); 11.22 (s, 1H, NH). 13C NMR (DMSO) δ: 25.1; 25.8; 46.2; 63.6; 82.3; 102.3; 124.3; 124.7; 125.6; 126.3; 126.7; 126.8; 126.9; 127.8; 128.5; 128.6; 130.4; 130.5; 140.0; 150.1; 162.6; 164.4. Anal. Calcd for C24H21O4N3 (415.43) C, 69.38; H, 5.10; N, 10.12. Found: C, 69.39; H, 5.05; N, 10.13 22Ab: (0.26 g, 93%) White solid from ethanol, mp > 200 °C. IR: νNH 3168, νCO 1720 cm−1. 1H NMR (DMSO) δ: 1.83 (s, 3H, CH3); 2.04 (s, 3H, CH3); 2.79 (m, 2H, CH2); 6.16 (t, 1H, J = 7 Hz, O-CHN); 7.58 (m, 4H, arom.); 7.81 (d, 1H, J = 7 Hz, N-CH); 7.86 (d, 1H, J = 9 Hz, arom.); 8.04 (m, 1H, arom.); 8.13 (m, 2H, arom.); 8.67 (s, 1H, arom.); 11.75 (s, 1H, NH). 13C NMR (DMSO) δ: 24.9; 25.7; 524

dx.doi.org/10.1021/jo302346a | J. Org. Chem. 2013, 78, 516−526

The Journal of Organic Chemistry

Article

(5) (a) Pistarà, V.; Corsaro, A.; Chiacchio, M. A.; Greco, G.; Quadrelli, P. ARKIVOC 2011, 270−285. (b) Chiacchio, U.; Genovese, F.; Iannazzo, D.; Piperno, A.; Quadrelli, P.; Corsaro, A.; Romeo, R.; Valveri, V.; Mastino, A. Bioorg. Med. Chem. 2004, 12, 3903−3909. (c) Chiacchio, U.; Genovese., F.; Iannazzo, D.; Librando, V.; Merino, P.; Rescifina, A.; Romeo, R.; Propopio, A.; Romeo, G. Tetrahedron 2004, 60, 441−448. (d) Saita, M. G.; Chiacchio, U.; Iannazzo, D.; Merino, P.; Piperno, A.; Previtera, T.; Rescifina, A.; Romeo, G.; Romeo, R. Nucleosides, Nucleotides Nucleic Acids 2003, 22, 739−742. (e) Chiacchio, U.; Corsaro, A.; Iannazzo, D.; Piperno, A.; Pistarà, V.; Rescifina, A.; Romeo, A.; Valveri, V.; Mastino, A.; Romeo, G. J. Med. Chem. 2003, 46, 3696−3702. (6) (a) Quadrelli, P.; Scrocchi, R.; Caramella, P.; Rescifina, A.; Piperno, A. Tetrahedron 2004, 60, 3643−3651. (b) Quadrelli, P.; Mella, M.; Carosso, S.; Bovio, B.; Caramella, P. Eur. J. Org. Chem. 2007, 6003−6015. (c) Quadrelli, P.; Mella, M.; Assanelli, G.; Piccanello, A. Tetrahedron 2008, 64, 7312−7317. (d) Savion, M.; Memeo, M. G.; Bovio, B.; Grazioso, G.; Legnani, L.; Quadrelli, P. Tetrahedron 2012, 68, 1845−1852. (7) (a) Quadrelli, P.; Mella, M.; Paganoni, P.; Caramella, P. Eur. J. Org. Chem. 2000, 2613−2620. (b) Quadrelli, P.; Fassardi, V.; Cardarelli, A.; Caramella, P. Eur. J. Org. Chem. 2002, 2058−2065. (8) For antiviral activity, the manuscript is in preparation. See also: (a) Kitade, Y.; Kojima, H.; Zulfiqur, F.; Kim, H. S.; Wataya, Y. Bioorg. Med. Chem. Lett. 2003, 13, 3963−3965. (b) Ramesh, N. G.; Klunder, A. J. H.; Zwanenburg, B. J. Org. Chem. 1999, 64, 3635−3641. (c) Wu, J.; Schneller, S. W.; Seley, K. L.; DeClerq, E. Heterocycles 1998, 47, 757−763. (d) Siddiqi, S. M.; Raissian, M.; Schneller, S. W.; Ikeda, S.; Snoeck, R.; Andrei, G.; Balzarini, J.; DeClerq, E. Bioorg. Med. Chem. Lett. 1993, 3, 663−666. (e) Da Silva, A. D.; Coimbra, E. S.; Fourrey, J. L.; Machado, A. S.; Robert-Gero, M. Tetrahedron Lett. 1993, 34, 6745−6748. (f) Koga, M.; Schneller, S. W. Tetrahedron Lett. 1990, 31, 5861−5864. (9) Moggio, Y.; Legnani, L.; Bovio, B.; Memeo, M. G.; Quadrelli, P. Tetrahedron 2012, 68, 1384−1392. (10) (a) Kirby, G. W. Chem. Soc. Rev. 1977, 6, 1−24. (b) Sato, T.; Aoyagi, S.; Kibayashi, C. Org. Lett. 2003, 5, 3839−3842. (c) Momiyama, N.; Yamamoto, H. Org. Lett. 2002, 4, 3579−3582. (d) Ware, R. W., Jr.; Day, C. S.; King, S. B. J. Org. Chem. 2002, 67, 6174−6180. (e) Chow, C. P.; Shea, K. J.; Sparks, S. M. Org. Lett. 2002, 4, 2637−2640. (f) Surman, M. D.; Mulvihill, M. J.; Miller, M. J. Tetrahedron Lett. 2002, 43, 1131−1134. (g) Surman, M. D.; Mulvihill, M. J.; Miller, M. J. J. Org. Chem. 2002, 67, 4115−4121. (h) Iwasa, S.; Tajima, K.; Tsushima, S.; Nishiyama, H. Tetrahedron Lett. 2001, 42, 5897−5899. (i) Flower, K. R.; Lightfoot, A. P.; Wan, H.; Whiting, A. J. Chem. Soc., Perkin Trans. 1 2002, 2058−2064. (j) Iwasa, S.; Fakhruddin, A.; Tsukamoto, Y.; Kameyama, M.; Nishiyama, H. Tetrahedron Lett. 2002, 43, 6159−6161. (k) Pulacchini, S.; Sibbons, K. F.; Shastri, K.; Motevalli, M.; Watkinson, M.; Wan, H.; Whiting, A.; Lightfoot, A. P. Dalton Trans. 2003, 2043−2052. (l) Adam, W.; Bottke, N.; Krebs, O.; Saha-Moller, C. R. Eur. J. Org. Chem. 1999, 1963−1965. (11) (a) Quadrelli, P.; Gamba Invernizzi, A.; Caramella, P. Tetrahedron Lett. 1996, 37, 1909−1996. (b) Quadrelli, P.; Mella, M.; Gamba Invernizzi, A.; Caramella, P. Tetrahedron 1999, 55, 10497− 10510. (12) Quadrelli, P.; Campari, G.; Mella, M.; Caramella, P. Tetrahedron Lett. 2000, 41, 2019−2022. (13) Quadrelli, P.; Mella, M.; Caramella, P. Tetrahedron Lett. 1999, 40, 797−800. (14) Quadrelli, P.; Mella, M.; Caramella, P. Tetrahedron Lett. 1998, 39, 3233−3236. (15) (a) Bodnar, B. S.; Miller, M. J. Angew. Chem., Int. Ed. 2011, 50, 5630−5647. (b) Miller, M. J. Proceedings of the International Symposium on Advances in Synthetic, Combinatorial, and Medicinal Chemistry, Moscow, Russia, May 5−8, 2004; T14 and P140. (c) Vogt, P. F.; Miller, M. J. Tetrahedron 1998, 54, 1317−1348. (d) Boger, L.; Weinreb, S. M. Hetero Diels-Alder Methodology in Organic Synthesis; Academic Press: San Diego, CA, 1987.

Table 2 Ar-CONO

alkene couples

8P

9d/e 0.86 9d/f 1.00 9d/e 0.77 9d/f 0.85

8M

ratios from data of Table1 9d/e 9d/e 9d/e 9d/e

0.77 0.87 0.67 0.75

program. Any IC90 value of ≥10 μg/mL is considered “Active” for antitubercular activity. For further information, see: http://www.taacf. org. X-ray Crystallography. A summary of crystal data, data collection, and structure refinement of compounds 11Md and 21A is given in Table S2 of the Supporting Information. The structures are solved by direct methods. Non-hydrogen atoms were refined anisotropically, and hydrogen atoms, located from the difference Fourier synthesis, were refined isotropically.40 CCDC 905532 for compound 11Md and CCDC 905533 for compound 21A contain the supplementary data. These data can be obtained free of charge at www.ccdc.cam.ac.uk/conts/retrieving.html [or from the Cambridge Crystallography Data Centre 12, Union Road, Cambridge CB2 IEZ, U.K., Fax: +44−1223/336033; E-mail: deposit@ ccdc.cam.ac.uk].41



ASSOCIATED CONTENT

S Supporting Information *

NMR spectra of new compounds, X-ray crystallographic data, and crystallographic data in CIF format. This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS Financial support by the University of Pavia, MIUR (PRIN 2008, CUP: F11J10000010001 and PRIN 2012, CUP: B11J12002450001) and EUTICALS S.p.a. - V.le Milano, 86/ 88- 26900 Lodi, Italy, is gratefully acknowledged. The authors wish to thank the Fondazione Banca del Monte di Lombardia for a research grant.



REFERENCES

(1) (a) Herdewijn, P. Drug Discovery Today 1997, 2, 235−242. (b) Wang, P.; Hong, J. H.; Cooperwood, J. S.; Chu, C. K. Antiviral Res. 1998, 40, 19−44. (c) Piperno, A.; Chiacchio, M. A.; Iannazzo, D.; Romeo, R. Curr. Med. Chem. 2006, 13, 3675−3695. (2) (a) De Clercq, E. Biochim. Biophys. Acta 2002, 1587, 258−275. (b) De Clercq, E. Med. Res. Rev. 2002, 22, 531−542. (c) Bricaud, A. D.; Herdewijn, P.; De Clercq, E. Biochem. Pharmacol. 1983, 3583− 3588. (3) (a) Wu, Q.; Simons, C. Synthesis 2004, 1533−1553 and references therein. (b) Rajappan, V. P.; Yin, X.; Schneller, S. W. Tetrahedron 2002, 58, 9889−9895. (c) Ishikura, M.; Murakami, A.; Katagiri, N. Org. Biomol. Chem. 2003, 1, 452−453. (d) Jeong, L. S.; Yoo, S. J.; Lee, K. M.; Koo, M. J.; Choi, W. J.; Kim, H. O.; Park, J. G.; Lee, S. K.; Chun, M. W. J. Med. Chem. 2003, 46, 201−202. (e) Gouverneur, V.; McCarthy, S. J.; Mineur, C.; Belotti, D.; Dive, G.; Ghosez, L. Tetrahedron 1998, 54, 10537−10554. (f) Kirby, G. W.; Nazeer, M. J. Chem. Soc., Perkin Trans. 1 1993, 1397−1402. (4) (a) Chang, C. N.; Doong, S. L.; Zhou, J. H.; Beach, J. W.; Jeong, L. S.; Chu, C. K.; Tsai, C. H.; Cheng, Y. C. J. Biol. Chem. 1992, 267, 13938−13942. (b) Grove, K. L.; Guo, X.; Lui, S. H.; Gao, Z.; Chu, C. K.; Cheng, Y. C. Cancer Res. 1995, 55, 3008−3011. 525

dx.doi.org/10.1021/jo302346a | J. Org. Chem. 2013, 78, 516−526

The Journal of Organic Chemistry

Article

(36) Fukuda, S.; Kamimura, A.; Kanemasa, S.; Hori, K. Tetrahedron 2000, 56, 1637−1647. (37) (a) Gonzalez, M. P.; Teran, C.; Fall, Y.; Teijeira, M.; Besada, P. Bioorg. Med. Chem. 2005, 13, 601−608. (b) Tchilibon, S.; Joshi, B. V.; Kim, S.-K.; Duong, H. T.; Gao, Z.-G.; Jacobson, K. A. J. Med. Chem. 2005, 48, 1745−1758. (c) Joshi, B. V.; Moon, H. R.; Fettinger, J. C.; Marquez, V. E.; Jacobson, K. A. J. Org. Chem. 2005, 70, 439−447. (38) Collins, L. A.; Franzblau, S. G. Antimicrob. Agents Chemother. 1997, 41, 1004−1009. (39) Grundmann, C.; Grünanger, P. The Nitrile Oxide; SpringerVerlag: Heidelberg, Germany, 1971. (40) North, A. C. T.; Phillips, D. C.; Mathews, F. S. Acta Crystallogr., Sect. A 1968, 24, 351−359. (41) (a) Sheldrick, G. M. SHELXL-93: Program for Crystal Structure Analysis; University of Gö ttingen: Gö ttingen, Germany, 1993. (b) Johnson, C. K. ORTEP Report ORNL-3793; Oak Ridge National Laboratory: Oak Ridge, TN, 1965.

(16) (a) Quadrelli, P.; Mella, M.; Piccanello, A.; Romano, S.; Caramella, P. J. Org. Chem. 2007, 72, 1807−1810. (b) Quadrelli, P.; Romano, S.; Piccanello, A.; Caramella, P. J. Org. Chem. 2009, 74, 2301−2310. (c) Adam, W.; Krebs, O. Chem. Rev. 2003, 103, 4131− 4146. (17) Johnston, B. D.; Czyzewska, E.; Oehlschlager, A. C. J. Org. Chem. 1987, 52, 3693−3697. (18) Kann, N.; Rein, T.; Akermark, B.; Helquist, P. J. Org. Chem. 1990, 55, 5312−5323. (19) For recent direct observation of acyl nitroso species, see: (a) Cohen, A. D.; Zeng, B.-B.; King, S. B.; Toscano, J. P. J. Am. Chem. Soc. 2003, 125, 1444−1445. (b) Evans, A. S.; Cohen, A. D.; GurardLevin, Z. A.; Kebede, N.; Celius, T. C.; Miceli, A. P.; Toscano, J. P. Can. J. Chem. 2011, 89, 130−138. (20) (a) Motorina, I. A.; Sviridova, L. A.; Golubeva, G. A.; Bundel, Y. G. Tetrahedron Lett. 1989, 30, 117−120. (b) Motorina, I. A.; Sviridova, L. A.; Golubeva, G. A.; Bundel, Y. G. Khim. Geterotsikl. Soedin. 1990, 976−979. (21) Grundmann, C.; Dean, J. M. J. Org. Chem. 1965, 30, 2809− 2812. (22) (a) Brochard, L.; Joseph, B.; Viaud, M.-C.; Rollin, P. Synth. Commun. 1994, 24, 1403−1414. (b) Da Ros, T.; Prato, M.; Lucchini, V. J. Org. Chem. 2000, 65, 4289−4297. (23) (a) Vorbrüggen, H.; Krolikiewicz, K.; Bennua-Skalmowski, B. U.S. Patent 5,750,676, 1998. (b) Vorbrüggen, H.; Krolikiewicz, K.; Bennua, B. Chem. Ber. 1981, 114, 1234−1255. (c) Vorbrüggen, H.; Höfle, G. Chem. Ber. 1981, 114, 1256−1268. (d) Vorbrüggen, H.; Bennua, B. Chem. Ber. 1981, 114, 1279−1286. (24) (a) Chiacchio, U.; Corsaro, A.; Iannazzo, D.; Piperno, A.; Pistarà, V.; Procopio, A.; Rescifina, A.; Romeo, G.; Romeo, R.; Siciliano, M. C. R.; Valveri, E. ARKIVOC 2002, 159−167. (b) Gotkowska, J.; Balzarini, J.; Piotrowska, D. G. Tetrahedron Lett. 2012, 53, 7097−7100. (25) For acetylation of 5-hydroxy-isoxazolidines, see: (a) Xiang, Y.; Gong, Y.; Zhao, K. Tetrahedon Lett. 1996, 37, 4877−4880. (b) Zhang, X.; Qing, F.-L; Yu, Y. J. Org. Chem. 2000, 65, 7075−7082. (c) Zhang, X.; Xia, H.; Dong, X.; Jin, J.; Meng, W.-D; Qing, F.-L. J. Org. Chem. 2003, 68, 9026−9033. (26) Yoshimura, Y.; Kitano, K.; Yamada, K.; Satoh, H.; Watanabe, M.; Miura, S.; Sakata, S.; Sasaki, T.; Matsuda, A. J. Org. Chem. 1997, 62, 3140−3152. (27) Caddell, J. M.; Chapman, A. M.; Cooley, B. E.; Downey, B. P.; LeBlanc, M. P.; Jackson, M. M.; O’Connell, T. M.; Phung, H.-M.; Roper, T. D.; Xie, S. J. Org. Chem. 2004, 69, 3112−3215. (28) (a) Chiacchio, U.; Saita, M. G.; Crispino, L.; Gumina, G.; Mangiafico, S.; Pistarà, V.; Romeo, G.; Piperno, A.; De Clercq, E. Tetrahedron 2006, 62, 1171−1181. (b) Chiacchio, U.; Iannazzo, D.; Piperno, A.; Romeo, R.; Romeo, G.; Rescifina, A.; Saglimbeni, M. Bioorg. Med. Chem. 2006, 14, 955−959. (29) (a) Shin, D.; Sinkeldam, R. W.; Tor, Y. J. Am. Chem. Soc. 2011, 133, 14912−14915. (b) Hernandez, A. R.; Kool, E. T. Org. Lett. 2011, 13, 676−679. (c) Wilson, J. N.; Kool, E. T. Org. Biomol. Chem. 2006, 4, 4265−4274. (30) Okamoto, A.; Tainaka, K.; Nishiza, K.; Saito, I. J. Am. Chem. Soc. 2005, 127, 13128−13129. (31) Charubala, R.; Maurinsh, J.; Rö sler, A.; Melguizo, M.; Jungmann, O.; Gottlieb, M.; Lehbauer, J.; Hawkins, M.; Pfleiderer, W. Nucleosides Nucleotides 1997, 16, 1369−1378. (32) (a) Leonard, N. J.; Sprecker, M. A.; Morrice, A. G. J. Am. Chem. Soc. 1976, 98, 3987−3994. (b) Hernandez, A. R.; Kool, E. T. Org. Lett. 2011, 13, 676−679. (33) (a) Wilhelmsson, L. M. Q. Rev. Biophys. 2010, 43, 159−183. (b) Hikida, Y.; Kimoto, M.; Yokoyama, S.; Hirao, I. Nat. Protoc. 2010, 5, 1312−1323. (c) Valis, L.; Wagenknecht, H.-A. Synlett 2005, 2281− 2284. (34) Sekine, M.; Oeda, Y.; Iijima, Y.; Taguchi, H.; Ohkubo, A.; Seio, K. Org. Biomol. Chem. 2011, 9, 210−218. (35) Bair, K. W.; Andrews, C. W.; Tuttle, R. L.; Knick, V. C.; Cory, M.; McKee, D. D. J. Med. Chem. 1991, 34, 1983−1990. 526

dx.doi.org/10.1021/jo302346a | J. Org. Chem. 2013, 78, 516−526